• Aucun résultat trouvé

Comparison of cardiovascular parameter estimation methods using swine data

N/A
N/A
Protected

Academic year: 2021

Partager "Comparison of cardiovascular parameter estimation methods using swine data"

Copied!
35
0
0

Texte intégral

(1)

The MIT Faculty has made this article openly available.

Please share

how this access benefits you. Your story matters.

Citation

Arai, Tatsuya et al., "Comparison of cardiovascular parameter

estimation methods using swine data." Journal of Clinical

Monitoring and Computing 34, 2 (April 2020): 261–70 ©2019 Authors

As Published

https://dx.doi.org/10.1007/s10877-019-00322-y

Publisher

Springer Netherlands

Version

Author's final manuscript

Citable link

https://hdl.handle.net/1721.1/129354

Terms of Use

Creative Commons Attribution-Noncommercial-Share Alike

(2)

Author accepted manuscript

Cite this article as: Tatsuya Arai, Kichang Lee and Richard J. Cohen, Comparison of car-diovascular parameter estimation methods using swine data, Journal of Clinical Monitoring and Computinghttps://doi.org/10.1007/s10877-019-00322-y

This Author Accepted Manuscript is a PDF file of an unedited peer-reviewed manuscript that has been accepted for publication but has not been copyedited or corrected. The official version of record that is published in the journal is kept up to date and so may therefore differ from this version.

Terms of use and reuse: academic research for non-commercial purposes, see here for full terms.https://www.springer.com/aam-terms-v1

(3)

Author accepted manuscript

Comparison of Cardiovascular Parameter Estimation Methods Using Swine

Data

Tatsuya Arai1, and Kichang Lee2,3,Richard J. Cohen2,

Department of Aeronautics and Astronautics1 and Institute for Medical Engineering and Science2 Massachusetts Institute of Technology, Cambridge, MA 02139; the Cardiovascular Research Center3,

Massachusetts General Hospital, Boston, MA 02114

Running head: Cardiovascular Parameter Estimation

Corresponding Author:

(4)

Author accepted manuscript

Institute for Medical Engineering and Science

Massachusetts Institute of Technology

45 Carleton Street, Room E25-324

Cambridge, MA 02142

Voice: (617) 452-2718

Fax: (617) 253-3019

(5)

Author accepted manuscript

ABSTRACT 1

In this study, new and existing methods of estimating stroke volume, cardiac output and total

2

peripheral resistance from analysis of the arterial blood pressure waveform were tested over a wide range

3

of conditions. These pulse contour analysis methods (PCMs) were applied to data obtained in six swine

4

during infusion of volume, phenylephrine, dobutamine, isoproterenol, esmolol and nitroglycerine as well

5

as during progressive hemorrhage. Performance of PCMs were compared using true end-ejection

6

pressures as well as estimated end-ejection pressures.

7

There was considerable overlap in the accuracies of the PCMs when using true end-ejection

8

measures. However, for perhaps the most clinically relevant condition, where radial artery pressure is the

9

input, only Wesseling’s Corrected Impedance method and the Kouchoukos Correction method achieved

10

statistically superior results.

11

We introduced a method of estimating end-ejection by determining when the systolic pressure

12

dropped to a value equal to the sum of the end-diastolic pressure plus a fraction of the pulse pressure.

13

The most accurate estimation of end-ejection was obtained when that fraction was set to 60% for the

14

central arterial pressure and to 50% for the femoral and radial arterial pressures.

15

When the estimated end-ejection measures were used for the PCMs that depend on end-ejection

16

measures, when radial artery pressure was used as the input, only Wesseling’s Corrected Impedance

17

method and the modified Herd’s method achieved statistically superior results.

18

This study provides a systematic comparison of multiple PCMs’ ability to estimate stroke volume,

19

cardiac output, and total peripheral resistance and introduces a new method of estimating end-systole.

20 21

Key words — Arterial blood pressure; cardiac output; stroke volume; total peripheral resistance; pulse

22

contour method

(6)

Author accepted manuscript

(7)

Author accepted manuscript

INTRODUCTION 25

When managing patients undergoing high-risk surgeries (i.e., liver transplantation) or in the setting of 26

an intensive care unit (ICU), monitoring cardiovascular hemodynamic information such as stroke volume 27

(SV), cardiac output (CO), and/or total peripheral resistance (TPR) is critically important. In general, 28

these parameters respond much more quickly to stresses (i.e., hemorrhage) than does arterial blood 29

pressure (ABP) which is continuously controlled by multiple physiological feedback and control 30

mechanisms to maintain a homeostatic state [1]. Thus, the ability to monitor SV or CO may enable 31

clinical intervention at an earlier stage prior to the development of hypotension, shock, and/or organ 32

damage during surgeries or ICU stays. 33

The most commonly accepted method to estimate CO in clinical settings is pulmonary artery 34

thermodilution, which involves injecting a bolus of cold liquid through a central venous catheter into the 35

right atrium and measuring the temperature change in the pulmonary artery [2, 3]. In general, 36

thermodilution requires pulmonary artery catheterization, which is associated with cardiovascular risks 37

such as carotid artery puncture (when accessing the internal jugular vein), cardiac arrhythmia, bleeding, 38

embolism, clotting, and infection [4, 5]. Transpulmonary thermodilution has become an alternative to 39

pulmonary artery thermodilution [6]. However, previous research has shown several limitations 40

associated with its use [7, 8]. 41

Even though continuous thermodilution CO measurement could provide a continuous trend of CO 42

[9], thermodilution method cannot continuously measure SV on a beat-to-beat basis and has significant 43

limitations [10, 11]. Therefore, many studies have thus been devoted to developing non-invasive or 44

minimally invasive methods to continuously estimate cardiovascular parameters. These methods include 45

Doppler ultrasound, transesophageal echocardiography, and impedance plethysomography [12-15]. 46

However, due to various reasons, such as lack of accuracy, not providing continuous measurement, 47

technical difficulties, requiring a medical specialist, and/or economic reasons, these systems are not 48

popularly used and/or used only for calibration purposes in the clinical setting. 49

(8)

Author accepted manuscript

Since the arterial pulse is readily accessible, it has been commonly used to estimate the 50

cardiovascular parameters. Specifically, mathematical analysis of the continuous ABP, termed a pulse 51

contour method (PCM), has been extensively studied to estimate cardiovascular parameters [16-25]. 52

However, the clinical use of this method has also been limited due to its inaccuracy. 53

The present study aimed to evaluate new algorithms to estimate continuous cardiovascular 54

hemodynamic parameters. These methods were validated with measured CO using the true “gold 55

standard for aortic blood flow (ABF) measurement method” – Transonic’s ultrasonic flow probe placed 56

on the aortic arch of the study animal – and these predictive accuracies were compared with existing PCM 57

algorithms. In addition, for a fair comparison, a new algorithm for beat-to-beat identification of arterial 58

end-ejection blood pressure from peripheral arteries was incorporated into the cardiovascular 59

hemodynamic parameter estimation methods. 60

61

METHODS 62

The algorithms described in this section were evaluated using previously reported data (21). The 63

following is a brief summary of the protocol. Six Yorkshire swine (30–34kg) were studied. The 64

experimental protocol conformed to the Guide for the Care and Use of Laboratory Animals and was 65

approved by the MIT Committee on Animal Care. The animals were pre-anesthetized with intramuscular 66

telazol, xylazine, and atropine prior to endotracheal intubation. The swine were then maintained in a deep 67

plane of anesthesia using inhaled anesthetic isoflurane (0.5-4 %), a mixture of oxygen and ambient air. 68

Positive-pressure mechanical ventilation at a rate of 10-15 breathes/min, and a tidal volume of 10 ml/kg 69

was employed. 70

Central ABP (CAP) was measured from the thoracic aorta using a micromanometer-tipped catheter 71

(SPC 350, Millar Instruments, Houston, TX). Femoral ABP (FAP) and radial ABP (RAP) were measured 72

using external fluid-filled pressure transducer (TSD104A, Biopac Systems, Santa Barbara, CA). The 73

(9)

Author accepted manuscript

chest was opened with a midline sternotomy. ABF was recoded using an ultrasonic flow probe placed 74

around the aortic root for reference CO (T206 with A-series probes, Transonic Systems, Ithaca, NY). 75

ABF, ECG, and ABPs were interfaced to a microcomputer via an analog-to-digital conversion system 76

(MP150WSW, Biopac Systems, Santa Barbara, CA) at a sampling rate of 250 Hz and 16-bit resolution. 77

In each animal, a subset of the following interventions was performed over the course of 75 to 150 78

min to vary the cardiac output and other hemodynamic parameters: infusions of volume, phenylephrine, 79

dobutamine, isoproterenol, esmolol, nitroglycerine, and progressive hemorrhage. To achieve substantial 80

cardiac output changes in a short period (15-20 mins), several infusion rates were implemented followed 81

by brief recovery periods (about 5 min). Also, hemorrhage was performed until a substantial change in 82

cardiac output was observed. At the conclusion of the experiment, the animal was euthanized with the 83

injection of sodium pentobarbital. 84

85

Algorithms 86

Modified Herd’s Method: Pulse pressure (PP) is the difference between systolic blood pressure (SBP)

87

and diastolic blood pressure (DBP) and is regarded as a proportional measure of SV [16]. The algorithm 88

is based on the Windkessel model with impulse ejection of SV [28]. The drawback of using PP as a 89

proportional measure of SV is the inaccuracy introduced because of the finite duration of ejection and the 90

distortion/alteration of the ABP waveform as it propagates through the arterial tree. In general, as the 91

ABP waveform propagates through the tapered and bifurcated peripheral arterial branches, the SBP 92

increases and the ABP waveform width becomes narrower. 93

To overcome this latter issue, Herd et al. used mean arterial pressure (MAP) instead of SBP, since 94

MAP is less sensitive to this distortion [17]. However, when MAP is calculated by averaging the ABP 95

waveform, the value of MAP can be affected by the duration of the diastolic interval, resulting in an SV 96

estimation error. For example, a longer diastolic interval would result in a smaller SV estimate – even 97

(10)

Author accepted manuscript

though diastole follows the completion of ejection, and thus the length of diastole cannot affect the value 98

of the preceding SV. To overcome this limitation, we used mean pressure during ejection instead of mean 99

pressure averaged over the entire beat: 100

(Equation 1)

101

= arterial compliance, = arterial blood pressure waveform, and = ejection period. DBP is

102

the end-diastolic blood pressure of the preceding beat. 103

CO was estimated from time-averaging the SV values and TPR was calculated using the following 104

equation (Ohm’s law): 105

TPR CO

MAP  (Equation 2)

106

CO and TPR estimates in the following methods were obtained in the same manner. 107

Auto-Regressive with Exogenous input (ARX) Model: We recently introduced a novel algorithm to 108

continuously estimate beat-to-beat ABF waveforms by analysis of the ABP signal. SV can be yielded by 109

the beat-to-beat integral of the ABF waveform, and CO can be calculated by the time average of ABF 110

over number of beats in a unit time. 111

In this section, the ABF estimation method will be briefly summarized (see Ref 26 for more details). 112

The mathematical model of the system can be described as an ARX input model that relates the ABP 113

values, , to the ABF values, : 114

(Equation 3)

115

where, are the autoregressive coefficients, is the parameter length, α is the weighting coefficient 116

for the exogenous input , and is noise. 117

(11)

Author accepted manuscript

Because the input ABF is approximately zero during diastole, the autoregressive coefficients can 118

be obtained by using a least-squares method to solve Equation 3: 119

(Equation 4)

120

where, designates a sample point during diastole. 121

The coefficients was obtained by solving the matrix equation using Matlab (Mathworks, Natick, 122

MA). A 17-beat moving window size was empirically found to be optimal with our algorithm for 123

estimating the coefficients and was therefore adopted. The autoregressive coefficient length was 124

chosen to minimize 125

The exogenous input weighting coefficient (α) was obtained by taking the average of both sides of 126

Equation 4: 127

(Equation 5)

128

where, MAP/CO can be obtained from Ohm’s law (Equation 2). 129

TPR is related to the and the characteristic time constant of the system (τ): 130

(Equation 6)

131

where, τ can be obtained by analyzing the terminal exponential decay curve of the impulse response of 132

the system : 133

(Equation 7)

134

Equations 5 and 6 can be combined to compute α: 135

(Equation 8)

136

Thus, instantaneous ABF can be expressed as: 137

(12)

Author accepted manuscript

(Equation 9)

138

The integral of was calculated on a beat-to-beat basis to obtain proportional SV estimates, and 139

the time average of over six minutes was calculated to obtain a proportional estimate of the CO 140

(proportionality constant being ). Thus, the algorithm presented here provides a comprehensive set of 141

proportional cardiovascular parameters (ABF, SV, CO, and TPR) based on an analysis of ABP 142

waveforms. 143

The calculated CO, SV, and TPR using these two methods were compared with those using the 144

previously reported methods. 145

Existing Pulse Contour Methods: Table 1 summarizes the existing cardiovascular parameter estimation

146

methods that were reported to be competitive in previous comparison studies [23, 27]. 147

Earlier works assumed that the arterial trees are represented by a two-parameter Windkessel model 148

accounting for the total compliance of the large arteries [arterial compliance ( )] and the TPR of small 149

arteries. During the diastolic period, the time constant (τ) is equal to the product of TPR and and the 150

proportional CO can be estimated using the time-averaged ABP and time constant [30]. Mukkamala et al. 151

calculated the time constant of the Windkessel model using an autoregressive moving average analysis 152

using arterial pressure and PP inputs to estimate the terminal projected exponential pressure decay during 153

diastole [21]. 154

Erlanger and Hooker described a relationship between SV and the PP suggesting that SV is 155

proportional to the PP [16]. Meanwhile, Herd et al. used MAP instead of SBP recorded in the ascending 156

aorta in the PP method to estimate robust SV [17]. When intra-aortic pressure is being measured 157

continuously, it is a relatively simple matter to subtract DBP from MAP and to multiply by the heart rate 158

(HR) to estimate CO. 159

(13)

Author accepted manuscript

Liljestrand-Zander reported that varied throughout the cardiac cycle and was dependent on ABP. 160

They used the inversely proportional relationship between and ABP to correct the non-linearity [20]. 161

Researchers also reported that SV is proportional to the area under the systolic region of the ABP 162

waveform [18, 19, 24, 25]. Kouchoukos et al. [19] and Wesseling et al. [25] proposed an empirical and 163

simple correction factor to the systolic area method to account for some source of error in ABP 164

fluctuations during the systolic period. Sun et al. [23] estimated SV using the root-mean-square of the 165

ABP waveform, which was claimed as one component of the LiDCOplus PulseCO method (LiDCO Ltd., 166

London, England). 167

The aforementioned methods use information regarding end-ejection. Traditionally, researchers have 168

used the dicrotic notch as an indicator of end-ejection. However, identifying the dicrotic notch can be 169

challenging since the dicrotic notch is often not detectable, particularly in the peripheral ABP signal. For 170

this reason, we estimated the end-systolic pressure values using the partial PP model. 171

Partial Pulse Pressure Model: An end-diastole always comes after a systolic peak. At end-ejection, the

172

pressure value is less than peak SBP. One can estimate the end-ejection pressure to correspond to the 173

ABP at the point in time when ABP falls to a value given by the following equation: 174

(Equation 10)

175

where, , , and are pressure values at end-ejection, end-diastole (previous beat), and peak

176

systole, respectively. 177

As examples, end-ejections identified by the 50% PP and 90% PP are shown in Figure 1. The time 178

stamp of can be regarded as the time of an estimated end-ejection. To determine the accuracy of the 179

PP model, we compared duration of diastole as estimated from the difference between the end-ejection 180

time determined by the partial PP Model and the onset of ejection as determined from the ABP signal 181

with the “true” duration of diastole as measured from the ABF signal. It was necessary to measure 182

(14)

Author accepted manuscript

duration of diastole because both end-ejection and onset of ejection time estimates in FAP and RAP are 183

delayed with respect to the true times of end-ejection and onset of ejection in the ABF signal measured in 184

the central aorta. We then determined the optimal value of the fraction f for each of the CAP, FAP and 185

RAP signals. The partial PP end-ejection identification method was then applied to the PCMs for 186

estimating SV, CO, and TPR. 187

The values of SV, CO, or TPR determined using the various algorithms are estimated to within a 188

proportionality constant (determined by ). Therefore, the comparison of estimated to measured values 189

of SV, CO, or TPR was achieved in each animal by adjusting the mean of each estimated parameter to 190

match the mean of the measured value. 191

For all methods, end-diastolic measures were computed from the preceding cardiac cycle. The 192

estimation errors are defined as root normalized mean squared error (RNMSE): 193

(Equation 11)

194

where, and are the measured and estimated values (i.e., SV, CO, and TPR), respectively, is

195

the number of data points, and is the number of free parameters. 196

RNMSEs of SV, CO, and TPR of each method with the true end-ejection pressure information were 197

compared with the other methods using analysis of variance (ANOVA). In addition, RNMSEs of SV, 198

CO, and TPR of each method with estimated end-ejection pressure using the partial PP model were 199

compared with the other six methods using ANOVA. If a significant difference was observed, simple 200

effects analysis with Duncan test was used to examine pair-wise differences (SAS 9.4). Statistical 201

significance was accepted at P<0.05. 202

RESULTS 203

(15)

Author accepted manuscript

Interventions resulted in a wide range of changes of CO (1.3 – 5.8 L/min), MAP (27 – 127 mmHg), 204

and HR (91 – 204 bpm). Table 2 summarizes the physiological ranges of the data sets. Over 68,000 205

beats were processed and analyzed for ABF and hemodynamic parameters. Figure 2 shows the SV, CO, 206

and TPR estimation errors with different methods. While there was considerable overlap in the 207

accuracies of the PCM estimates, for perhaps the most clinically relevant estimations, which use RAP as 208

the input, only the Wesseling’s Corrected Impedance and the Kouchoukos Correction methods achieved 209

statistically superior results for all three of the estimated hemodynamic parameters. 210

Using the partial PP model, the end-ejection identification errors were minimum when the fraction f 211

in Eq. 10 was set to 60% for CAP, 50% for FAP, and 50% for RAP - as shown in Table 3. Thus, the most 212

accurate estimation of end-ejection was obtained when end-ejection was estimated to occur when systolic 213

pressure dropped to a value equal to the end-diastolic pressure plus 60% (50%) of the PP for the CAP (for 214

the FAP and RAP). Here the diastolic pressure and PP were referenced to the previous beat end-215

diastolic pressure. 216

In Figure 3, we show the RNSME results when using the estimated end-ejection time and pressures 217

for methods that depend on the end-ejection measures. The above optimal values of f were used here. 218

For the most clinically relevant condition where RAP is the input, only Wesseling’s Corrected Impedance 219

method and the modified Herd’s method achieved statistically superior results for all three of the 220

estimated hemodynamic parameters. In particular, Wesseling’s Corrected Impedance method provided 221

the lowest RNSMEs of 15.7% (SV), 12.3% (CO) and 12.9% (TPR). 222

223

DISCUSSION 224

In this paper, new algorithms were tested to estimate cardiovascular hemodynamic information. An 225

algorithm using the ARX model to continuously estimate ABF by the analysis of peripheral ABP 226

waveform was used to calculate CO, SV, and TPR. In addition, the modified Herd’s method was tested 227

(16)

Author accepted manuscript

and systemically compared with the existing hemodynamic parameter estimation methods using the same 228

ABP dataset. We also tested existing PCM algorithms and evaluated the impact of estimating end-229

ejection time and pressure on the performance of the PCM algorithms. 230

There was considerable overlap in the accuracies of the PCM estimates when using true end-ejection 231

pressures. However, for perhaps the most clinically relevant estimations, which use radial artery pressure 232

as the input, only the Wesseling’s Corrected Impedance and the Kouchoukos Correction methods 233

achieved statistically superior results for all three of the estimated hemodynamic parameters. 234

All the methods incorporate their own assumptions in cardiovascular physiology. Cardiovascular 235

hemodynamic parameter estimation methods need to work under a wide set of physiological conditions in 236

clinical and research settings. The parameters of Wesseling’s Corrected Impedance method [25] were 237

empirically obtained from a human study. In this method, the systolic area under the ABP curve above 238

DBP was scaled using a scaling factor that is a function of HR and MAP. Although the scaling factor 239

formula was obtained from healthy male subjects in their twenties, the method achieved low errors when 240

applied to the swine data sets, indicating that the human and swine cardiovascular system may be similar 241

in terms of applicability of the model. The Kouchoukos Correction method [19] includes a simple 242

correction factor (TS/TD) to model run-off blood flow during systole. Although the correction factors are

243

in both cases empirical, the Wesseling’s and Kouchoukos’s methods achieved lower errors than several 244

theoretical model-based methods. 245

Liljestrand-Zander’s method [20] unexpectedly generated high errors with the swine data, although it 246

has been reported to have the best agreement with the thermodilution CO in ICU patient data sets [23]. 247

This could be attributed to the nature of the ICU data sets. Because clinicians attempt to maintain the 248

patient’s ABP and CO, there is less variation in these signals obtained from patients than those obtained 249

during animal experiments in which these signals can be varied more widely using a variety of 250

interventions. Thus, methods that tend to provide stable estimates may appear to perform better with 251

(17)

Author accepted manuscript

patient data where the majority of the input parameters are stable. However, the utility of a method to 252

measure CO and other cardiac hemodynamic parameters is to identify those rare occasions when these 253

parameters deviate substantially from their normal values. The data analysis employed in this study was 254

designed to weigh the tail values specifically to test this aspect. 255

For the hemodynamic parameter estimation, end-systole (onset of diastole) needs to be determined for 256

each beat. In practical settings, a standard method to detect the end-systole (onset of diastole) is the use 257

of the dicrotic notch in ABP waveforms. However, the dicrotic notch does not always exist in the ABP 258

waveform. Therefore, we evaluated the performance of a model to estimate end-ejection (Eq. 10). 259

The most accurate estimation of end-ejection was obtained when end-ejection was estimated to occur 260

when systolic pressure dropped to a value equal to the end-diastolic pressure plus 60% (50%) of the PP 261

for the CAP (for the FAP and RAP). Here the end-diastolic pressure and PP are referenced to the 262

previous beat end-diastolic pressure. 263

In Figure 3, we show the RNSME results when using the estimated end-ejection pressures for 264

methods that depend on the end-ejection pressure. Here, for the most clinically relevant condition when 265

radial artery pressure is the input, only Wesseling’s Corrected Impedance method and the modified 266

Herd’s method achieved statistically superior results for all three of the estimated hemodynamic 267

parameters. In particular, Wesseling’s Corrected Impedance method provided the lowest RNSMEs. 268

The ARX algorithm utilizes the notion that the input to the arterial system is zero during diastole. In 269

the ABF estimation routine, 17 diastolic ABP waveforms were used to obtain the autoregressive (AR) 270

parameter and the AR parameters were integrated into the ARX model and applied to the entire ABP 271

waveform to obtain the ABF waveform. The AR parameters were also used to obtain the characteristic 272

time constant as well as the scaling factor to properly scale the estimated ABF. The 17-beat moving 273

window size was empirically chosen. If the window is too short, one cannot excite enough modes to 274

identify the system. On the other hand, if the window is too long, one cannot assume time-invariance of 275

(18)

Author accepted manuscript

the pertinent cardiovascular system. This method provides not only proportional SV, CO, and TPR, but 276

also instantaneous ABF waveforms without training data sets or demographic hemodynamic parameters - 277

arguably one of the most comprehensive estimation algorithms to our knowledge. 278

The classical Windkessel model assumes exponential decay during diastole and this model can be 279

described as a low-order AR model. The present ARX algorithm, on the other hand, obtains higher-order 280

AR parameter from diastolic ABP waveforms. The advantage of the present ARX algorithm is that it 281

appears to take into account possible distortion in the diastolic ABP waveforms in that the filter created 282

by the algorithm can reliably reconstruct the systolic ABF waveform. The distortion property may vary 283

from artery to artery, as well as from subject to subject. The algorithm could obtain individual parameters 284

unique to each arterial line of each subject on a beat-to-beat basis. 285

Further development of accurate end-systole identification methods (e.g., perhaps incorporating heart 286

sounds) might lead to more robust SV, CO, and TPR estimation using the new methods. Future work is 287

needed to apply and validate the algorithm with abnormal beats, such as premature beats and in heart 288

failure models. The methods could also be applied to optimizing SV when programming atrioventricular 289

time delay for conventional pacemakers and timing parameters for biventricular pacing. 290

One limitation of the current work is that the animal data involved using healthy pigs (~35 kg) with 291

normal hearts. Further studies would be necessary to apply the methods described here under a variety of 292

pathological clinical conditions (e.g. heart failure). The methods described here also need to be evaluated 293

using human data under various clinical conditions and populations. 294

295

CONCLUSION 296

(19)

Author accepted manuscript

This paper tested new algorithms to estimate hemodynamic parameters (SV, CO, and TPR) by 297

analysis of the ABP signal. Additionally, a new algorithm to identify end-ejection was implemented in 298

conventional and the new hemodynamic parameter estimation algorithms. 299

There was considerable overlap in the accuracies of the PCM estimates when using true end-ejection 300

pressures. However, for perhaps the most clinically relevant estimations, which use radial artery pressure 301

as the input, only the Wesseling’s Corrected Impedance and the Kouchoukos Correction methods 302

achieved statistically superior results for all three of the estimated hemodynamic parameters. 303

The Wesseling’s Corrected Impedance method and the modified Herd’s method performed best 304

among methods that depended on end-ejection time or pressure when estimated, rather than true, values 305

of end-ejection measures were used. In particular, the Wesseling’s Corrected Impedance method 306

provided the lowest errors. 307

(20)

Author accepted manuscript

Compliance with ethical standards 309

310

Conflict of Interest Richard Cohen is a co-inventor on two patents in the area of hemodynamic 311

parameter estimation assigned to the Massachusetts Institute of Technology (MIT) which have been 312

licensed to Retia Medical, LLC. Dr. Cohen is not otherwise involved with the company. The other 313

authors declare no conflicts. 314

315

Ethical approval All procedures performed in studies involving animals were in accordance with the 316

ethical standards of the institution. 317

318 319 320

(21)

Author accepted manuscript

REFERENCES 321

[1] H. Barcroft, O.G. Edholm, J. McMichael, and E.P. Sharpey-Schafer, "Posthaemorrhagic fainting. 322

Study by cardiac output and forearm flow," Lancet, vol. 1, pp. 489-491, 1944. 323

[2] W. Ganz, R. Donoso, H.S. Marcus, J.S. Forrester, and H.J. Swan, "A new technique for 324

measurement of cardiac output by thermodilution in man," Am J Cardiol, vol. 27, pp. 392-6, 1971. 325

[3] W. Ganz and H.J. Swan, "Measurement of blood flow by thermodilution," Am J Cardiol, vol. 29, 326

pp. 241-6, 1972. 327

[4] G.R. Manecke, Jr., J.C. Brown, A.A. Landau, D.P. Kapelanski, C.M. St Laurent, and W.R. Auger, 328

"An unusual case of pulmonary artery catheter malfunction," Anesth Analg, vol. 95, pp. 302-4, 329

2002. 330

[5] J.S. Vender and H.C. Gilbert, “Monitoring the anesthetized patient,” 3 ed. Philadelphia: Lippincott-331

Raven Publishers, 1997. 332

[6] X. Monnet and J.L. Teboul, “Transpulmonary thermodilution: advantages and limits,” Crit Care, 333

vol. 21, pp. 147, 2017. 334

[7] L. Huter, K.R. Schwarzkopf, N.P. Schubert, and T. Schreiber, “The level of cardiac output affects 335

the relationship and agreement between pulmonary artery and transpulmonary aortic thermodilution 336

measurements in an animal model,” J Cardiotheorac Vasc Anesth, vol. 21, pp. 659-63, 2007. 337

[8] K. Staier, M. Wilhelm, C. Wiesenack, M. Thoma, and C. Keyl, “Pulmonary artery vs. 338

transpulmonary thermodilution for the assessment of cardiac output in mitral regurgitation: a 339

prospective observational study,” Eur J Anaesthesiol, vol. 29, pp. 431-7, 2012. 340

[9] F.G. Mihm, A. Gettinger, C.W. Hanson, 3rd, H.C. Gilbert, E.P. Stover, J.S. Vender, B. Beerle, and 341

G. Haddow, "A multicenter evaluation of a new continuous cardiac output pulmonary artery 342

catheter system," Crit Care Med, vol. 26, pp. 1346-50, 1998. 343

(22)

Author accepted manuscript

[10] M. Botero, D. Kirby, E.B. Lobato, E.D. Staples, and N. Gravenstein, "Measurement of cardiac 344

output before and after cardiopulmonary bypass: Comparison among aortic transit-time ultrasound, 345

thermodilution, and noninvasive partial CO2 rebreathing," J Cardiothorac Vasc Anesth, vol. 18, pp. 346

563-72, 2004. 347

[11] W.H. Fares, S.K. Blanchard, G.A. Stouffer, P.P. Chang, W.D. Rosamond, H.J. Ford, and R.M. Aris, 348

"Thermodilution and Fick cardiac outputs differ: impact on pulmonary hypertension evaluation," 349

Can Respir J, vol. 19, pp. 261-6, 2012. 350

[12] M.E. Blohm, D. Obrecht, J. Hartwich, G.C. Mueller, J.F. Kersten, J. Weil, and D. Singer, 351

"Impedance cardiography (electrical velocimetry) and transthoracic echocardiography for non-352

invasive cardiac output monitoring in pediatric intensive care patients: a prospective single-center 353

observational study," Crit Care, vol. 18, pp. 603, 2014. 354

[13] S. Schubert, T. Schmitz, M. Weiss, N. Nagdyman, M. Huebler, V. Alexi-Meskishvili, F. Berger, 355

and B. Stiller, "Continuous, non-invasive techniques to determine cardiac output in children after 356

cardiac surgery: evaluation of transesophageal Doppler and electric velocimetry," J Clin Monit 357

Comput, vol. 22, pp. 299-307, 2008. 358

[14] S. Scolletta, F. Franchi, S. Romagnoli, R. Carla, A. Donati, L.P. Fabbri, F. Forfori, J.M. Alonso-359

Inigo, S. Laviola, V. Mangani, G. Maj, G. Martinelli, L. Mirabella, A. Morelli, P. Persona, D. 360

Payen, and PulseCOval Group, "Comparison Between Doppler-Echocardiography and Uncalibrated 361

Pulse Contour Method for Cardiac Output Measurement: A Multicenter Observational Study," Crit 362

Care Med, vol. 44, pp. 1370-9, 2016. 363

[15] R.P. Patterson, "Fundamentals of impedance cardiography," IEEE Eng Med Biol Mag, vol. 8, pp. 364

35-8, 1989. 365

(23)

Author accepted manuscript

[16] J. Erlanger and D.R. Hooker, "An experimental study of blood-pressure and of pulse-pressure in

366

man," Johns Hopkins Hosp Rep, vol. 12, pp. 145-378, 1904.

367

[17] J.A. Herd, N.R. Leclair, and W. Simon, "Arterial pressure pulse contours during hemorrhage in 368

anesthetized dogs," J Appl Physiol, vol. 21, pp. 1864-8, 1966. 369

[18] W.B. Jones, L.L. Hefner, W.H. Bancroft, Jr., and W. Klip, "Velocity of blood flow and stroke 370

volume obtained from the pressure pulse," J Clin Invest, vol. 38, pp. 2087-90, 1959. 371

[19] N.T. Kouchoukos, L.C. Sheppard, and D.A. McDonald, "Estimation of stroke volume in the dog by 372

a pulse contour method," Circ Res, vol. 26, pp. 611-23, 1970. 373

[20] G. Liljestrand and E. Zander, "Vergleichende Bestimmung des Minutenvolumens des Herzens beim 374

Menschen mittels der Stickoxydulmethode und durch Blutdruckmessung," Zeitschrift fur die 375

gesamte experimentelle Medizin, vol. 59, pp. 105-122, 1928. 376

[21] R. Mukkamala, A.T. Reisner, H.M. Hojman, R.G. Mark, and R.J. Cohen, "Continuous cardiac 377

output monitoring by peripheral blood pressure waveform analysis," IEEE Trans Biomed Eng, vol. 378

53, pp. 459-67, 2006. 379

[22] T. Parlikar, T. Heldt, G.V. Ranade, and G. Verghese, "Model-Based Estimation of Cardiac Output 380

and Total Peripheral Resistance," presented at the Computers in Cardiology, pp. 379-82, 2007. 381

[23] J.X. Sun, A.T. Reisner, M. Saeed, T. Heldt, and R.G. Mark, "The cardiac output from blood 382

pressure algorithms trial," Crit Care Med, vol. 37, pp. 72-80, 2009. 383

[24] P.D. Verdouw, J. Beaune, J. Roelandt, and P.G. Hugenholtz, "Stroke volume from central aortic 384

pressure? A critical assessment of the various formulae as to their clinical value," Basic Res 385

Cardiol, vol. 70, pp. 377-89, 1975. 386

(24)

Author accepted manuscript

[25] K.H. Wesseling, B. De Werr, J.A.P. Weber, and N.T. Smith, "A simple device for the continuous 387

measurement of cardiac output. Its model basis and experimental verification," Adv Cardiovasc 388

Phys, vol. 5, pp. 16-52, 1983. 389

[26] T. Arai, K. Lee, R.P. Marini, and R.J. Cohen, "Estimation of changes in instantaneous aortic blood 390

flow by the analysis of arterial blood pressure," J Appl Physiol, vol. 112, pp. 1832-8, 2012. 391

[27] T. Parlikar, "Modeling and Monitoring of Cardiovascular Dynamics for Patients in Critical Care," 392

Department of Electrical Engineering and Computer Science, Massachusetts Institute of 393

Technology, Cambridge, 2007. 394

[28] M.J. Bourgeois, B.K. Gilbert, G. Von Bernuth, and E.H. Wood, “Continuous determination of beat 395

to beat stroke volume from aortic pulse pressures in the dog,” Circ Res, vol. 39, pp. 15-24, 1976. 396 397 398 399 400 401 402

(25)

Author accepted manuscript

Fig 1. End-systole (ejection) defined by means of the partial pulse pressure (PP). 50% and 90 % PP are 403 shown as examples. 404 405 406 407 408

(26)

Author accepted manuscript

409 410 a) SV 411 412 413 b) CO 414

(27)

Author accepted manuscript

415

c) TPR 416

Figure 2. The SV, CO, and TPR estimation errors with different methods using the measured end-417

ejection pressure. 418

* P<0.05 lower than other methods with central arterial pressure (CAP). † P<0.05 lower than other 419

methods with femoral arterial pressure (FAP). ‡ P<0.05 lower than other methods with radial arterial 420

pressure (RAP) 421

MH: modified Herd's method; WCIM: Wesseling’s corrected impedance method; Kou: Kouchoukos 422

correction; Wind: Windkessel model AUC_D: area under the curve with end-diastolic ABP value 423

subtracted; AUC: area under the systolic curve; ACP: alternating current power; PP: pulse pressure; Herd: 424

Herd’s pulse pressure; BBA: beat-to-beat average; ARMA: autoregressive moving average; Lilj: 425

Liljestrand-Zander’s. 426

(28)

Author accepted manuscript

428 a) SV 429 430 431 b) CO 432 433

(29)

Author accepted manuscript

434

c) TPR 435

Figure 3. The SV, CO, and TPR estimation errors with different methods using the partial pulse 436

pressure model to estimate end-ejection pressure. 437

438

* P<0.05 lower than other methods with central arterial pressure (CAP). † P<0.05 lower than other 439

methods with femoral arterial pressure (FAP). ‡ P<0.05 lower than other methods with radial arterial 440

pressure (RAP) 441

ARX, ARX model with exogenous input; MH: modified Herd's method; WCIM: Wesseling’s corrected 442

impedance method; Kou: Kouchoukos correction; Wind: Windkessel model AUC_D: area under the 443

curve with end-diastolic ABP value subtracted; AUC: area under the systolic curve. 444 445 446 447 448

(30)

Author accepted manuscript

Table 1.Existing cardiovascular hemodynamic parameter estimation methods. 449

450

Windkessel Model [28] Pulse Pressure [16]

Herd’s Pulse Pressure [17] Liljestrand-Zander’s [20]

Beat-to-Beat Average (BBA) Model [22] Systolic Area [18], [24]

Wesseling’s Corrected Impedance [25]

Kouchoukos Correction [19]

Alternating Current Power [23]

Auto-Regressive Moving Average [21]

(31)

Author accepted manuscript

451

PP: pulse pressure; SBP: systolic blood pressure; DBP: diastolic blood pressure; MAP: mean arterial 452

pressure; Ca: compliance of the arterial tree; CO: cardiac output; SV: stroke volume; F: aortic blood

453

flow; T: duration of cardiac cycle; P1: arterial blood pressure at the beginning of the beat; P2: arterial

454

blood pressure at the end of the beat; τ: time constant of arterial system; P: arterial blood pressure; t: 455

time; HR: heart rate; TS: systolic duration in Kouchoukos correction method; TD: diastolic duration in

456

Kouchoukos method; tED: time at which end-diastole occurs; tEE: time at which end-ejection occurs; 457

: autoregression coefficients; : moving average coefficients; TPR: total peripheral resistance. 458

459 460

(32)

Author accepted manuscript

Table 2. Summary of hemodynamic parameters (Mean ± SD) of the six swine data sets. 461

462 463

CO: cardiac output, SV: stroke volume, FAP: femoral arterial pressure, RAP: radial arterial pressure, HR: 464 heart rate. 465 466 CO (L/min) SV (mL) FAP (mmHg) RAP (mmHg) HR (bpm) 1 3.6±1.0 28.4±5.8 63±19 61±19 129±29 2 3.2±0.6 25.0±5.0 83±21 73±20 135±38 3 4.0±0.7 31.7±7.1 83±16 87±15 133±32 4 3.2±0.6 25.2±4.3 89±19 79±18 129±34 5 3.3±0.5 26.7±6.4 80±21 85±29 130±32 6 3.4±1.2 28.5±8.1 72±19 75±20 130±26 Mean 3.5±0.8 27.5±6.7 79±21 76±21 131±32

(33)

Author accepted manuscript

Table 3. Summary of diastolic interval error ± SD (%). 467

468

CAP FAP RAP

40% PP 17.1 ± 11.6 5.9 ± 8.0 6.0 ± 14.2 50% PP 10.1 ± 9.0 -3.3 ± 5.5 -1.4 ± 12.5 60% PP 1.8 ± 6.9 -7.4 ± 4.1 -10.8 ± 6.8 70% PP -4.7 ± 4.3 -10.0 ± 3.9 -13.8 ± 7.1 80% PP -8.2 ± 3.5 -12.8 ± 3.9 -17.1 ± 8.2 90% PP -11.3 ± 3.8 -16.3 ± 4.3 -23.8 ± 11.3 469

PP: pulse pressure, CAP: central arterial pressure, FAP: femoral arterial pressure, RAP: radial arterial 470

pressure. 471

(34)

Author accepted manuscript

GLOSSARY: 473

ABF: aortic blood flow 474

ABP: arterial blood pressure 475

ACP: alternating current power 476

AR: autoregressive 477

ARMA: autoregressive moving-average model 478

ARX: autoregressive with exogenous input 479

AUC: area under the systolic 480

AUC_D: Area under the curve with end-diastolic ABP value subtracted 481

BBA: beat-to-beat averaged model 482

Ca: arterial compliance

483

CAP: central arterial pressure 484

CO: cardiac output 485

DBP: diastolic blood pressure 486

FAP: femoral arterial pressure 487

HR: heart rate 488

ICU: intensive care unit 489

MAP: mean arterial pressure 490

(35)

Author accepted manuscript

MH: modified Herd's method 491

PCM: pulse contour method 492

PP: Pulse pressure 493

RAP: radial arterial pressure 494

RNMSE: root normalized mean squared error 495

SBP: systolic blood pressure 496

SV: stroke volume 497

TPR: total peripheral resistance 498

WCIM: Wesseling’s corrected impedance method 499

500 501

Figure

Figure 2.  The SV, CO, and TPR estimation errors with different methods using the measured end-end-417
Table 1. Existing cardiovascular hemodynamic parameter estimation methods.
Table 2. Summary of hemodynamic parameters (Mean ± SD) of the six swine data sets.

Références

Documents relatifs

Simos, Jesu’s Vigo Aguiar, A modified Runge-Kutta method with phase- lag of order infinity for the numerical solution of the Schr¨ odinger equation and related problems, Comput.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Abstract— This work proposes an optimal approach for parameter estimation in a landslide motion, based on the so-called adjoint method. The system is described by an

A new method is presented for the measurement of the mean lepton lifetime using events in which 's are pair-produced and both 's decay to hadrons and. Based on the

The Multi-dimensional Optimal Order Detection (MOOD) method has been designed by authors in [5] and extended in [7] to reach Very-High-Order of accuracy for systems of Conservation

I MPACT OF THE L OADS ON T HE A CCES I MPEDANCE The measurement of the access impedance when the customer installation is connected to the electrical network is

The formulation, termed Modified Wave Finite Ele- ment (MWFE), allows the global wave modes of a given multi-layered system to be constructed from the local wave mode bases attached

On peut faire l’hypothèse que la littérarité est en jeu aussi dans un cas comme celui du blog de poèmes de David Nadaud, Sur le coin d’une table (Figure 3) : la sociologie de